Exploring the Acid Gas Sorption Properties of Oxidatively Degraded

Dec 26, 2018 - Abstract Image. Amine-supported mesoporous oxide materials have proven to be effective acid gas sorbents. While the primary application...
1 downloads 0 Views 1MB Size
Article pubs.acs.org/EF

Cite This: Energy Fuels XXXX, XXX, XXX−XXX

Exploring the Acid Gas Sorption Properties of Oxidatively Degraded Supported Amine Sorbents Matthew E. Potter,†,§ Kyeong Min Cho,†,‡ Jason J. Lee,† and Christopher W. Jones*,† †

Downloaded via EASTERN KENTUCKY UNIV on January 11, 2019 at 19:10:57 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

School of Chemical & Biomolecular Engineering, Georgia Institute of Technology, 311 Ferst Drive, NW, Atlanta, Georgia 30332, United States ‡ Department of Chemical and Biomolecular Engineering (BK-21 plus), Korea Advanced Institute of Science and Technology, Daejeon 34141, Republic of Korea S Supporting Information *

ABSTRACT: Amine-supported mesoporous oxide materials have proven to be effective acid gas sorbents. While the primary application of these supported amine species has been CO2 capture, they have also shown to be proficient at adsorbing other damaging flue gas impurities such as SOx and NOx. The precise nature of the amine (primary, secondary, or tertiary) is known to dictate the gas−amine interactions, with tertiary amines of particular interest due to their inability to adsorb dry CO2, favoring SOx and NOx species. The different amine sites also provoke differences in oxidative stability, when exposed to temperatures similar to those used for thermal desorption procedures. Herein we focus on the structural and chemical changes that occur in a range of class 1 (amine-impregnated) and class 2 (amine-grafted) sorbents upon oxidation and correlate these with their variation in acid gas (CO2, NO2, and SO2) uptakes, as a function of the oxidation temperature. These studies suggest that oxidatively degraded or “spent” supported amine materials may have possible uses as NOx or SOx sorbents. Specifically, despite oxidative degradation these aminopolymer species maintain a reasonable level of NO2 uptake, despite losing the ability to capture SO2 or CO2, offering unique possibilities in selective NO2 capture.



INTRODUCTION Presently, fossil fuels are the main contributor to the global energy supply, resulting in the accelerated release of carbon dioxide and increased quantities of anthropogenic CO2. The concentration of atmospheric CO2 has received significant interest in recent decades, due to its strong correlation with environmental issues, such as global warming and numerous health issues, such as respiratory illnesses.1,2 As such, there is growing interest in the fields of carbon capture, utilization, and sequestration to control CO2 levels.3,4 The most evolved and widely implemented technology for CO2 separation from dilute gas streams utilizes liquid amines, such as aqueous monoethanolamine (MEA), which is the prototypical benchmark−capture system for postcombustion, or “flue-gas” scrubbing systems.5 While liquid amine technology is well understood and extensively employed, its use nonetheless presents a challenge. The small amine molecules are volatile and reactive, and as such, a significant quantity of the sorbent is lost throughout the process. Additionally, the solutions are highly corrosive, requiring specialized equipment.6 These issues could be circumvented by employing a solid sorbent.7−9 To date many solid systems, primarily porous materials, have been evaluated as potential CO 2 sorbents, including carbons,10,11 zeolites,12 metal organic frameworks,13,14 and supported amine materials.15−18 Supported amine materials have been of great interest as they can capture CO2 under both flue-gas (∼5−15 vol % CO2) and direct-air capture (400 ppm of CO2) conditions.19,20 This is further supported by their lower regeneration costs accompanied by high amine loadings that create a high density of sites for CO2 adsorption.21,22 © XXXX American Chemical Society

The use of fossil fuels also results in the production of SOx and NOx emissions, typically emissions from coal result in concentrations between 0.2−0.25 vol % and 0.15−0.2 vol % of flue gas, respectively.23,24 While the quantity of these species is much less than CO2, they can hinder carbon capture processes, with parts per million (ppm) levels of SO2 and NO2 in particular, resulting in the deactivation of amine-based sorbents via irreversible sorption.25 While selective catalytic reduction is commonly used to lower NOx emissions, it does not completely remove all of the NOx present. In light of this, processes to selectively remove these species have been implemented, such as scrubbing sulfur gases via CaCO3.26 However, this limestone process requires temperatures in excess of 800 °C to desorb the SO2 and, in doing so, forms a considerable amount of waste inorganic byproducts.27 As such, one could develop a process using supported amines, tailored for the selective reversible adsorption of SOx and NOx species to prescrub flue gas sources before carbon capture occurs.28 Supported amines have been grouped into four different classes depending on the nature and incorporation method of the amine species. Class 1 “molecular basket” species are the result of oligomeric or polymeric amine species, with varying distributions of primary, secondary, or tertiary amines residing within a mesoporous support (typically silica, alumina, or carbon).29−31 Examples include poly(allylamine) (PAA, primary amines only), linear poly(ethylenimine) (L-PEI, secondary amine rich), or branched poly(ethylenimine) (BReceived: October 21, 2018 Revised: December 22, 2018 Published: December 26, 2018 A

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels PEI, typically 44% primary, 33% secondary and 22% tertiary amines). These materials have received considerable attention, as they permit high amine loadings (5−10 mmol/g) for improved uptake and use of higher molecular weight polymers can limit amine evaporation or leaching.32 In contrast, class 2 species have a chemical bond between the support and the amine.33−35 This is typically achieved by anchoring a trialkoxysilane species, such as 3-aminopropyl triethoxysilane to the support, through Si−O−Si (C3H6NR2) bonds. These species are more resistant to chemical leaching; however, the grafting leads to lower amine loadings (typically ∼1−2 mmol/ g), limiting their ability to capture vast quantities of CO2. These silanes are relatively expensive as well. Classes 3 and 4 materials can be viewed as combination of the above two classes. Class 3 is the result of in situ polymerization, resulting in an anchored aminopolymer,36 whereas Class 4, which contains both anchored species and impregnated polyamine species within the same structure.37 In all cases, it is possible tune the nature of the alkylamine by selecting a specific polymer or organosilane.38 The contrasting behavior of primary, secondary, and tertiary amines is of great interest in optimizing selective sorption of acid gases.39−42 It is understood that under “dry” conditions, primary amines make the more effective supported amine CO2 sorbents, due to their ability to form strongly bound alkyl ammonium carbamate ionic pairs.39−41 Yet, under anhydrous conditions tertiary amines cannot form these ionic pairs due to the lack of N−H bonds. This means a third species, typically water, is required, leading to the formation of an alkylammonium bicarbonate species.42 As such, significant interest has developed in utilizing tertiary amines under dry conditions, to selectively adsorb SO2.43 SO2 has been shown to have a different binding mechanism than CO2, whereby a charge-transfer complex is formed by tertiary amines binding to SO2; R3N−SO2.44 Unlike CO2 sorption, recent literature suggests that secondary amines are more effective than tertiary or primary amines for SO2 sorption, but the reasons for this are currently not well-understood.44 Similarly the binding mechanism of NO2 has been hypothesized, but never confirmed, with studies showing that it is able to indiscriminately bind to all types of amines, with similar affinities.45 Aside from the differing adsorption behavior, another pivotal factor in the nature of the amine groups is their stability, specifically their reactivity with oxygen, at temperatures necessary for acid-gas desorption.46−50 It has been reported that secondary amines are the more reactive isolated amine species toward oxygen.47−50 In work by Sayari et al., secondary amine (N-methylaminepropyl silane, MAPS) groups were found to form imines, amides and nitrone species, through 13C MAS NMR, upon exposure to carbon-free air at 120 °C for 40 h.34 In contrast, no significant change was seen for the primary (3-aminopropyl silane, APS) or tertiary (N,N-dimethylaminepropyl silane, DMAPS) amine groups. It is also noted that while isolated primary amine groups were not susceptible to oxidation, they became less stable when in close proximity to secondary amines.47−50 L-PEI (with only secondary amines) and B-PEI (44% primary, 33% secondary, and 22% tertiary amines) were compared before and after exposure to oxidizing conditions (110 °C in air for 24 h). It was found that the combination of different amine sites in B-PEI resulted in lower stability than the (almost) purely secondary-amine L-PEI species, as the CO2 capacity of B-PEI had dropped by 98% after oxidation, whereas the L-PEI had only dropped by 54%.50

Once oxidatively degraded it is challenging to fully regenerate the supported amine sorbents, meaning they would typically be disposed for reprocessing. However, if alternative uses could be found for these sorbents then it would mean they could be “recycled” at the end of their carbon capture lifetime, making them highly attractive for large scale applications.51 To this end, there has seen a growing interest in propylene linked aminopolymers and dendrimers, which have been reported to show better oxidative stability than ethylene linked polymers during CO2 adsorption under oxidizing conditions.51,52 While the influence of oxidative degradation has been well studied on many supported amine species relevant to CO2 capture,51 very few reports exist on how this can influence the sorption properties of different amines for “amine-poisoning” gases such as SO2 and NO2. As such, in this work, we compare the influence of isolated primary, secondary, and tertiary class 2 aminosilane species, anchored on mesoporous SBA-15, for SO2, NO2, and CO2 (under both simulated flue-gas and atmospheric conditions) using dry gases to probe the lifetime of such species for selective acid gas sorption. Similarly we compare these findings with that of B-PEI and fully methylated B-PEI impregnated in SBA-15 to understand how different distributions of amine groups can lead to tailored acid-gas sorption and their resilience to oxidative degradation.



EXPERIMENTAL DETAILS

Chemicals. All chemicals were obtained from Sigma-Aldrich and used without further purification. SBA-15 Synthesis. SBA-15 was synthesized as per our previous work.15 First, 24.01 g of Pluronic P123 was initially dissolved in 630 mL of H2O and 120 mL of concentrated hydrochloric acid (37 wt %) while rapidly stirring at 40 °C to give a clear solution with white foam. Next, 52.68 g of tetraethyl orthosilicate was added dropwise, after which the system was left to stir for 20 h at 40 °C. The system was then aged under static conditions at 100 °C for 24 h. The mixture was then filtered and washed with 6 L of deionized water. The white powder was subsequently dried at 75 °C overnight prior to calcination. The powder was calcined by heating to 200 °C at a rate of 1.2 °C/min, holding for 1 h, and then increasing the temperature to 550 °C at a rate of 1.2 °C/min and holding for 4 h to yield a white powder. Methylating PEI. Branched PEI (800 MW) was methylated following the procedure by Sayari et al. First, 1 g of PEI was added to a solution of 12 mL of formaldehyde, 14 mL of formic acid, and 0.4 mL of water. The mixture was heated to 120 °C overnight under a nitrogen atmosphere. When cooled, the solvent was removed with a rotary evaporator at 45 °C. This mixture was made basic with 30 mL of a 20 wt % NaOH solution. KOH pellets were then added to the solution until the product was obtained as an oily organic layer. The methylated-PEI was then extracted with diethyl ether and dried over anhydrous Na2SO4. The orange product was then dried overnight at 110 °C for 12 h at 10 mbar. The purity of the compound was confirmed through 1H solution phase NMR using a Bruker DSX-400, with CDCl3 solvent. PEI Impregnation (Class 1 Samples). Initially, 1.0 g of the above SBA-15 was stirred with 15 mL of methanol at room temperature for 1 h. A 20 mL solution of methanol containing either 500 mg of 800 MW branched poly(ethylenimine) (PEI), or 700 mg of methylated poly(ethylenimine) (Me-PEI) was added to the slurry and stirred for a further 24 h at room temperature. The solvent was then removed under reduced vacuum at 50 °C to yield a white powder. The sample was then dried at 110 °C for 12 h at 10 mbar. Samples are labeled as MePEI-SBA or PEI-SBA accordingly. Amine-Grafting on Silica (Class 2 Samples). First, 1.0 g of the SBA-15 described above was dried overnight at 110 °C under 20 mbar of vacuum. This sample was then mixed in a solution containing 20 mL of dry toluene, 50 μL of deionized water, and 3 mmol of either B

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Figure 1. (A) N2 physisorption isotherms comparing SBA, PEI-SBA, and MePEI-SBA, showing a significant N2 uptake on impregnation of the aminopolymers. (B) Variation in organic content after oxidation treatment as a function of oxidation temperature of the PEI-SBA and MePEI-SBA series. 13 C MAS NMR. 13C cross-polarization magic angle spinning (CPMAS) solid-state nuclear magnetic resonance (NMR) measurements were carried out on a Bruker DSX-300 spectrometer. Samples (Class 1 and Class 2) were spun at a frequency of 10 kHz, and 18 000 scans were taken. FTIR. FTIR measurements were run on a Harrick DRIFTS on a Nicolet is10 with MCT detector for 32 scans and a 4 cm−1 resolution. The samples were activated on a high vacuum line at 110 °C for 3 h prior to measurement.

3-aminopropyl triethoxysilane (APS), 3-methylaminopropyl trimethoxysilane (MAPS) or N,N-dimethylaminopropyl trimethoxysilane (DMAPS). The solution was heated to 70 °C for 24 h and, then, filtered with 200 mL of toluene, hexane, and ethanol. The resulting powder was then dried overnight at 110 °C under 10 mbar of vacuum. Samples are labeled as APS/MAPS/DMAPS-SBA accordingly. Organic Content. The organic content of the Class 1, Class 2, and bare silicate samples was estimated using a Netzsch STA409PG TGA. Weight loss was calculated between 120 and 900 °C, under a combined flow of 90 mL/min of air and a 30 mL/min of nitrogen. The data were collected between 25 and 900 °C heating at a rate of 10 °C/min. Physisorption. Nitrogen physisorption was performed on a Micrometrics Tristar 3020 instrument at 77 K. Samples were degassed for 12 h at 110 °C prior to analysis for the Class 1, Class 2, and bare silica samples. Surface area was calculated using the BET model. Pore volumes and pore-size distributions were calculated using the Bdb-FHH model.53 Acid Gas Capacity Measurements. SO2, NO2, and CO2 capacities for the Class 1, Class 2, and bare silica samples were measured using a TA Instruments Q500 TGA. Samples were pretreated under a 100 mL/min flow of helium at 110 °C for 3 h. The sorbent uptake was then measured from the gain in mass after subsequent exposure to 10% CO2 in helium, 400 ppm of CO2 in helium, 200 ppm of SO2 in helium, or 200 ppm of NO2 in helium as appropriate, at 35 °C, flowed at 90 mL/min for 6 h, or until saturation, with a 10 mL/min balance helium flow. The setting 35 °C was chosen as the adsorption temperature for all gases for consistency and to allow for comparisons with our previous work.23 For NO2 desorption measurements, the adsorbed samples were exposed to a 100 mL/min flow of helium and heated to 110 °C for 3 h. The amount of desorbed species was quantified from the mass lost during this period. Sample Degradation. To evaluate the oxidative stability of the materials described above, 500 mg of the adsorbent was packed into a Pyrex tube, 1 cm in diameter, with a frit at the center to allow for flow of gas through the sample without loss of the adsorbent. To remove residual water from the system, the bed was first heated to 110 °C and nitrogen was passed through the bed at 20 mL/min for 1 h. The temperature was then set to the desired oxidation temperature, and the flow was switched to air oxygen at 20 mL/min. The air flow was stopped after 24 h, and the material was recovered for postoxidation chemical analysis.



RESULTS AND DISCUSSION Physicochemical Properties of Class 1 Sorbents. The integrity of the bare SBA-15 (referred to here on as SBA) support was confirmed using nitrogen physisorption (Figure 1 and S1). This showed a surface area of 793 m2 g−1 and a pore volume of 1.00 cm3 g−1; a narrow pore distribution was observed using the BdB-FHH method, showing an average pore diameter of 81 Å, all consistent with the previous literature.15,19,20 SBA-15 was then impregnated with both branched 800 MW PEI (henceforth known as PEI-SBA) and fully methylated branched 800 MW PEI (henceforth known as MePEI-SBA). The latter was synthesized as stated above, and its phase purity was confirmed through 1H NMR (Figure S2), showing excellent agreement with previous work by Sayari et al.24 When discussing the physicochemical properties, we present our findings per gram of SBA (gSBA−1), to normalize the sorbent to equivalent quantities to allow more insightful structural comparisons. In both cases the average pore size for the three samples (SBA, PEI-SBA, and MePEI-SBA) shows a significant decrease (from 81 to 63 Å) (Figure 1 and S1). A dramatic decrease in surface area and pore volume were also seen on impregnation, confirming that the polymers are occupying the mesopores of SBA. It is noted that the addition of MePEI decreases the surface area and pore volume more than PEI. This is partially attributed to the greater organic content of MePEI-SBA (0.70 g gSBA−1) over PEI-SBA (0.49 g gSBA−1) (Figure S1B), which was necessary to ensure approximately equivalent amounts of amines in the two systems. Here MePEI-SBA has an amine content of 12.2 mmol gSBA−1, and PEI-SBA has an amine content of 11.4 mmol gSBA−1. This level of PEI loading was purposefully chosen to C

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Table 1. Summarizing the Physicochemical Properties of the Bare Support and of Both Series of Class 1, Amine-Impregnated Sorbents, both Fresh and after Oxidation Treatment organic content/(g gSBA−1)

surface area/(m2 gSBA−1)

pore volume/(cm3 gSBA−1)

pore distribution/Å

average pore diameter/Å

SBA

N/A

793

1.00

70−102

81

PEI-SBA PEI90-SBA PEI110-SBA PEI130-SBA

0.49 0.42 0.34 0.42

305 267 382 345

0.65 0.60 0.77 0.70

53−70 53−75 58−77 58−72

63 63 71 64

MePEI-SBA MePEI90-SBA MePEI110-SBA MePEI130-SBA

0.70 0.67 0.39 0.52

148 151 269 234

0.36 0.38 0.57 0.54

58−78 53−67 57−74 57−75

64 58 63 63

system

Figure 2. N2 physisorption isotherms of (A) PEI impregnated SBA-15 (PEI-SBA) and (B) methylated PEI impregnated SBA-15 (MePEI-SBA), after oxidative treatment at the specified temperatures.

oxidation at 90 °C, is followed by a more rapid decrease at 110 °C. Previous work suggests that the primary modes of oxidative degradation are the formation of imines, amides, and conjugated amides.30 Of these, only imine formation reduces the organic content. Yet this does not occur at tertiary amines, and even quantitative imine formation in PEI-SBA cannot account for all the organic content lost. Therefore, it is likely that the decrease in organic content is due to aminopolymer evaporation, leaching, or cracking.32 A greater proportion of the organic content is lost in MePEI-SBA (45% from fresh to 110 °C) than PEI-SBA (31% from fresh to 110 °C, Table 1). Recent SANS and theoretical work suggests that initially aminopolymers form strong interactions with the pore wall through primary amine groups, creating a monolayer along the pore walls.19,39 After the monolayer formation, polymer− polymer interactions dominate as the polymer forms “plugs” in the mesopores, similar to the bulk (nonencapsulated) polymer.19,39 This helps rationalize why a greater decrease is observed in the MePEI-SBA, than PEI-SBA. As MePEI-SBA has no primary amines, the polymer−support interactions will be much weaker, leading to a greater amount of polymer mobility and potential loss. In both cases an increase in organic content is seen from 110 to 130 °C, which is likely due to oxygen addition to the aminopolymers, or fragments thereof, in forming amide or other oxygen-containing species above 110 °C.30 This is shown through FTIR analysis, where PEI110SBA, PEI130-SBA, MePEI110-SBA, and MePEI130-SBA all

achieve reasonably high amine loadings for acid gas capture, without completely filling the mesopores with PEI, as this leads to pore blockage which hinders the diffusion of gases through the polymer, lowering the amine efficiency.31 However, this difference may also be due to a difference in topology between the two systems.19,39 It is understood that PEI binds to SBA-15 through hydrogen-bonding interactions between silanol species and primary amines.39 As there are no primary amines in MePEI, the hydrogen bonding interactions will be more limited. Also, methylating all the amines likely makes the system more hydrophobic than materials made with conventional PEI. The combination of these two factors will likely result in MePEI adopting a different morphology to PEI in the pores, contributing to the physicochemical differences observed. The PEI-SBA and MePEI-SBA systems were oxidized in a controlled environment, to mimic oxidative degeneration of the sorbents. By simulating increased temperatures and air exposure, we sought insights into sorbent lifetime and stability on desorption. Both systems were oxidized for 24 h at 90, 110, and 130 °C to create a series of sorbents, labeled as (MePEI/ PEI)(90/110/130)-SBA to indicate the original sorbent and the oxidation temperature. In all cases, the organic content was monitored via TGA, both before and after the oxidation treatment (Figure S3). Both the PEI-SBA and MePEI-SBA series show similar trends (Figure S3). A gentle decrease in organic content, after D

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels show a noticeable peak in the region of 1670 cm−1, attributed to imine or amide formation (Figure S4). The peak is noticeably broad due to the overlap with physisorbed water at 1630 cm−1. A similar trend is also seen in the physisorption data (Figure 2 and S5), as both the pore volume and surface area of the 110 °C treated samples are significantly higher than the fresh systems (Table 1). This is in line with the theory that the pores are now less occluded due to loss of aminopolymer. Going from the 110 °C oxidation to 130 °C, in both systems, a reduction in both pore volume and surface area is observed, likely due to the increased molecular weight of the oxidized aminopolymers. We note subtle shifts in average pore diameter with increasing temperatures for the oxidation process, with larger pores reappearing in both the 110 and 130 °C treated materials, likely due to aminopolymer leaching. However, the PEI90-SBA shows a slightly higher pore volume and surface area than the fresh PEI-SBA. This subtle feature is likely due to the rearrangement of the polymer inside the pores, given that the organic content is similar (Table 1). It is, however, also possible that nondegraded PEI is leaching out, and the organic content is being artificially increased by the initial degradation of PEI. The chemical environment of both of the fresh aminopolymer systems and the systems oxidized at 130 °C were probed using 13C magic angle spinning solid state NMR (Figure 3). The fresh PEI-SBA and MePEI-SBA samples

peaks appear in the 180 to 150 ppm and 150 to 115 ppm ranges in both PEI130-SBA and MePEI130-SBA. The former set of peaks have been previously attributed to CO stretches from amide groups,30,33,34,50 showing that oxygen has clearly reacted with the amines present, in both species. The peaks at 150−115 ppm are attributed to CC and C−N bonds from heterocycles and aromatic species.33 It is interesting to note that these species were not seen by Sayari et al. using their solvent extraction method.50 However, such species may have coked onto the surface or were not separated during the solvent extraction protocol used. In this study we cannot clearly resolve any signals that are readily attributed to imine species (∼45 ppm), as these would overlap with the signals from the sp3 carbons in the fresh aminopolymers.50 However, the main signal in PEI-SBA shifts from 50 to 45 ppm on oxidation, which could suggest imine formation. This is in line with recent reports on degraded PEI,54 which showed the emergence of signals around 45 ppm in 13C NMR on oxidative degradation at 130 °C after 6 h in a 15% CO2 in air mixture. A less dramatic change is observed in this region or the MePEI-SBA systems, as imine species are unable to form from tertiary amines (Figure 3). However, there is still significant evidence for the formation of oxidative products in the MePEISBA system, showing significant degradation and amide formation.51 Physicochemical Properties of Class 2 Sorbents. To observe the influence of different amines (primary, secondary, and tertiary) in isolation, a range of class 2 sorbent species was also prepared and oxidized at 130 °C. The fresh samples were prepared to ensure they had similar amine loadings (2.4−2.7 mmol gSBA−1). In all cases, the fresh class 2 sorbents (APS/MAPS/DMAPSSBA) showed a lower surface area and pore volume than the bare SBA support (Table 2, Figure S6 and S7), indicating that the organosilane moieties had been successfully anchored into the mesopores. As a result, the average pore diameter was in all cases reduced from the 81 Å in the bare SBA to 71 ± 1 Å in the class 2 species. After oxidation at 130 °C the organic content of the class 2 species showed minor variations of ±10%. However, this is expected, as the modest amine loadings used means the mass resolution is not sufficient to conclusively determine whether the species has oxidized or not. As expected, no significant amine loss or leaching was observed, due to the strong chemical bond anchoring the organosilane to the surface. Nitrogen physisorption shows that the oxidation of the class 2 sorbents at 130 °C results in notable decreases in both pore volume and surface area (Table 2, Figure S6). This is attributed to subtle chemical transformation of the amines, as previously seen,46−50 which modify surface−amine interactions, changing the morphology of the

Figure 3. Solid state magic angle spinning cross-polarization 13C NMR of both fresh aminopolymers and after oxidation treatment at 130 °C.

exclusively show peaks in the 35 to 70 ppm region, representing aliphatic amine species.33,34,50,52 On oxidation,

Table 2. Summarizing the Physicochemical Properties of the Bare Support and the Three Different Class 2 Sorbents, Both before and after Oxidation Treatment at 130 °C system

organic content/(g gSBA−1)

surface area/(m2 gSBA−1)

pore volume/(cm3 gSBA−1)

pore distribution/Å

average pore diameter/Å

SBA APS-SBA APS130-SBA MAPS-SBA MAPS130-SBA DMAPS-SBA DMAPS130-SBA

N/A 0.12 0.13 0.16 0.15 0.17 0.16

793 391 286 470 408 459 439

1.00 0.81 0.62 0.76 0.74 0.95 0.76

70−102 64−83 65−85 63−86 63−88 64−86 63−79

81 71 72 72 71 70 70

E

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

concentrations of CO2 is therefore attributed specifically to the leaching, deactivation, or rearrangement of “bulk” aminopolymer, as seen by TGA (Figure S3, Table 1), leading to less amine sites available to bind 400 ppm of CO2. It is possible that PEI degradation is occurring at this temperature, but a more thorough study would be required to verify this hypothesis. Given the rapid change in CO2 uptake between 90 and 110 °C, it is likely the degradation occurs in this regime. Increasing the oxidation temperature to 110 °C results in a dramatic decrease in CO2 capacity at both CO2 concentrations (Figure 4, Table S1) and is similar to the bare SBA support, suggesting that the aminopolymer is now significantly less effective at adsorbing CO2. This rapid decline in performance is attributed not only to the aminopolymer leaching but also the oxidation of the amine species into amides and heterocycles, as seen by the 13C solid state NMR (Figure 3) and suggested by the theoretical amine efficiencies (Table S1).30,33,34,50 This is likely initiated by the secondary amines being in close proximity to other amines, as previously discussed.46−50 PEI130-SBA shows the same uptake as the bare SBA support. This suggests that almost all of the amines have been oxidized, to the point where they cannot chemisorb CO2. This trend mirrors previous work54 which showed that after just 6 h of oxidative degeneration, the CO2 uptake of PEI-SiO2 can decrease by 75%.54 For the Class 2 sorbents we see a clear trend whereby the primary amines (APS) are superior CO2 sorbents compared to secondary amines (MAPS) for both 400 ppm and 10 vol % CO2 feedstocks (Figure 5, Table S1).15−20 Isolated tertiary

tethered amines, possibly by forcing them off the pore walls. We note that we did not perform analogous 13C NMR studies of the oxidatively degraded class 2 materials, as this is already well documented.33 Acid-Gas Adsorption. The supported amine sorbents were tested for the uptake of various acid gases, including 10 vol % CO2 (mimicking flue gas), 400 ppm of CO2 (mimicking atmospheric CO2), 200 ppm of SO2, and 200 ppm of NO2. The latter two gases were chosen to evaluate the potential for using supported amines for impurity adsorption.25,43−45 In this section we normalize our results per gram of sorbent (g−1); this is due to the difficulties in comparing systems of different classes, with different amine loadings. Thus, this format allows for a more reasonable comparison of the uptake of the sorbents. The MePEI-SBA series showed no notable CO2 uptake at 35 °C under dry conditions (as expected); however PEI-SBA and PEI90-SBA both showed significant uptake, with fresh PEISBA showing amine efficiencies of 0.21 and 0.12 for the 10 vol % and 400 ppm of CO2 feeds, respectively (Table S1, Figure 4). The reduction in CO2 uptake with more dilute feedstocks is

Figure 4. Variation in CO2 uptake of the PEI impregnated SBA-15 series as a variation of oxidation treatment temperature for both 10 vol % and 400 ppm of CO2 feedstocks, compared to the bare SBA-15 support. Measurements were performed at 35 °C, flown at 90 mL/ min for 6 h, with a 10 mL/min balance helium flow.

likely due to less frequent CO2−amine interactions.19,20 Further, fewer sites can initiate the stronger CO2−amine interactions required to selectively capture CO2 at low concentrations (>60 kJ mol−1).15,40 The CO2 uptake using 400 ppm of CO2 for PEI-SBA and PEI90-SBA decreased by 29% (from 0.71 mmol g−1 to 0.52 mmol g−1); however, the organic content only decreased by 15% (from 0.48 to 0.41 g gSBA−1). Given this disparity, it is likely that the aminopolymer lost, or degraded, was the more effective species for adsorbing 400 ppm of CO2, which we previously suggested, through microcalorimetry, was the “bulk” or “plug” PEI molecules.15,19 In contrast, the CO2 uptake for 10 vol % CO2 decreased by roughly 16% (from 1.20 to 1.00 mmol g−1), in line with the organic content lost (Table 1). The higher concentration of CO2 is less challenging to adsorb, and therefore, a wider range of amine sites (both monolayer and “bulk” aminopolymer) can likely capture it.19,20 The lower uptake of PEI90-SBA for both

Figure 5. Showing the variation in CO2 uptake of class 2 sorbents, both fresh and after 130 °C oxidation treatment, compared with the bare SBA-15 support. Measurements were performed at 35 °C, flowed at 90 mL/min with either 400 ppm of CO2, or 10 vol % CO2 (as labeled) for 6 h, with a 10 mL/min balance helium flow.

amine (DMAPS) species showed little CO2 uptake for either feedstock, again due to the “dry” conditions. Once oxidized at 130 °C the MAPS-SBA system loses a significant amount of its CO2 uptake for the 10 vol % feedstock. In contrast the primary (APS) system seems more resilient to oxidation at 130 °C, with the uptake decreasing only slightly for the 10 vol % CO2. F

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Figure 6. (A) Influence of oxidation temperature on SO2 uptake with SBA impregnated with either PEI or methylated PEI, compared with the uptake of the bare SBA support. (B) Variation in SO2 uptake of Class 2 sorbents, both fresh and after 130 °C oxidation treatment, compared with the bare SBA-15 support. All measurements were performed at 35 °C, in gas flow at 90 mL/min with 200 ppm of SO2 until saturation, with a 10 mL/min balance helium flow.

tertiary amines present are able to adsorb the SO2 better than the mixture of primary/secondary/tertiary amines in the PEISBA. This result shows that while the MePEI system cannot adsorb CO2 in dry conditions, it is still equally as effective at adsorbing dry SO2 as branched PEI. Therefore, one could satisfactorily use MePEI-SBA as a system to selectively adsorb dry SO2.24 As the oxidation temperature increases to 90 °C, the SO2 uptake of both the MePEI-SBA and PEI-SBA systems decrease (Figure 6A, Table S2), with the MePEI-SBA SO2 uptake decreasing at a faster rate. This may be due to the greater aminopolymer loss in MePEI-SBA, likely through evaporation or cracking, as seen through TGA (Figure S3 and Table 1). Again the two sorbents show very similar SO2 uptakes (PEI90SBA, 1.44 mmol g−1; MePEI-90-SBA, 1.29 mmol g−1). On oxidation at 110 °C very little uptake is seen, again due to the oxidation of the amines, in both aminopolymers. This suggests there is little difference in stability between the fully methylated MePEI and the standard branched PEI for degradation by oxidation at this temperature. At 130 °C both systems have the same capacity as the SBA support, suggesting that few of the oxidized amine sites can chemisorb SO2. Comparing this with the Class 2 species (Figure 6B, Table S2) shows that the complete deactivation of the amines is not necessarily a strong function of proximity, as, once oxidized, the APS130-SBA and MAPS130-SBA systems show a similar level of SO2 uptake to SBA. However, DMAPS-SBA and DMAPS130-SBA show similar SO2 uptakes, far in excess of SBA. This suggests that while primary and secondary amines do deactivate in isolation, isolated tertiary amines are more resilient and do not degrade as readily, unlike tertiary amines in close proximity, such as MePEI. Again, the fresh sorbents follow the literature precedent that secondary amines are best suited to adsorbing SO2. However, this is also accompanied by more facile oxidative deactivation.46−50 NO2 sorption shows different trends to both CO2 and SO2. NO2 is known to adsorb more readily onto supported amines

This is in good agreement with the previous literature, whereby isolated secondary amines are more readily oxidized, with isolated primary amines being comparatively stable.46−50 Neither the MAPS-SBA or MAPS130-SBA species are effective for adsorbing 400 ppm of CO2, as they show similar uptakes to the bare SBA. However, for 400 ppm of CO2 the APS system shows a notable decrease in CO2 capacity when oxidized at 130 °C. This is likely because any site that can effectively interact with CO2 at such low concentrations will require a high binding energy (in the region of 60−80 kJ mol−1).15,40 However, these sites that interact strongly with CO2 appear to also react with oxygen under these conditions. Overall, these findings show that after oxidation at 130 °C, the Class 2 materials can still uptake significant amounts of 10 vol % CO2, confirming that the isolated amines are generally more stable than aminopolymers under the employed degradation conditions.50 This is most likely due to a lack of diamine structures and/or amine reorganization in Class 2 materials. While CO2 adsorption is a good probe for the stability of certain amines, SO2 and NO2 can probe a wider range of available amine sites.23,24,43−45 SO2 is known to favor secondary amine sites,23,24 whereas NO2 binds almost indiscriminately to amines.23 The combination of these probe molecules, and CO2, will provide a more holistic view of the impact of oxidative degradation within these aminopolymers. The oxidized PEI-SBA and MePEI-SBA series shows very similar behavior in the uptake of SO2 (Figure 6A, Table S2). Both show a gentle decrease in uptake after oxidation at 90 °C and, then, a rapid decrease in uptake after oxidation at 110 °C. Both systems show very similar uptake to the bare SBA system after oxidation at 130 °C. The initial SO2 uptake is marginally greater in the MePEI-SBA system (1.99 mmol g−1) than the PEI-SBA system (1.88 mmol g−1). This is likely due to the slightly higher amine loading in the MePEI-SBA system than the PEI-SBA system (12.1 and 11.3 mmol gSBA−1, respectively). This translates into a much higher amine efficiency for MePEISBA (0.54 versus 0.33 for PEI-SBA, Table S2), suggesting the G

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Figure 7. (A) Influence of the temperature of the oxidation treatment on NO2 uptake with SBA impregnated with either PEI or methylated PEI, compared with the uptake of bare SBA support. (B) Variation in NO2 uptake of Class 2 sorbents, both fresh and after 130 °C oxidation treatment, compared with the bare SBA-15 support. All measurements were performed at 35 °C, flow at 90 mL/min with 200 ppm of NO2 until saturation, with a 10 mL/min balance helium flow.

NO2 adsorption (Figure 7A, Table S2). It has already been established that very few unoxidized amines remain after oxidation at 130 °C, as seen by the loss of SO2 and CO2 uptake. Given this, we conclude that the oxidatively degraded amines (likely amides, imines, etc.) in the aminopolymers, are active NO2 sorbents, despite not being able to adsorb significant amounts of CO2 or SO2. This suggests that such species could be highly selective NO2 sorbents and have potential in NO2 capture. This could allow “used” (degraded) PEI-based sorbents that can no longer adsorb CO2 to be recycled for effective NO2 scrubbing. The behavior of the isolated class 2 species suggest that tertiary amines have a higher equilibrium capacity for NO2 uptake (Figure 7B, Table S2). However, the more revealing fact is that the APS130-SBA and DMAPS130-SBA show only a slight deviation from the NO2 uptake values of the fresh systems. In contrast, the MAPS130-SBA shows a significant decrease from the fresh system. Therefore, while NO2 is less selective about the type of amine it binds to, the oxidation products of isolated secondary amines, specifically, are not as effective for NO2 binding. However, the oxidation products from tertiary and primary amines are still able to effectively adsorb NO2. It is therefore likely that the more noticeable reduction in NO2 uptake in PEI-SBA over MePEI-SBA is due to the oxidized secondary amines in PEI-SBA being less effective for NO2 capture. However, it should be noted that in all cases the degraded Class 2 sorbents still significantly outperformed the bare SBA support, and therefore, the oxidized secondary amines are only comparatively less effective than the other supported amine species tested here. The influence of oxidation temperature on the desorption of NO2 was also studied (Table S3). We note that the PEI-SBA series desorbs roughly constant amounts of NO2 (57−67%) despite the range of oxidation temperatures studied. Similarly, there is little difference in the various Class 2 species on oxidation. We note, however, that the oxidation temperature greatly influences the MePEI-SBA series, as both the fresh MePEI-SBA and MePEI90-SBA desorb roughly 15−20% of

than both SO2 and CO2, hence larger uptake values are seen (Figure 7A, Table S2).23 It can also be seen that the vast majority of the NO2 uptake comes from the aminopolymer and not the SBA support, given the NO2 uptake values of the fresh aminopolymers (PEI-SBA 6.26 mmol g−1 and MePEI-SBA 4.88 mmol g−1) are significantly higher than the bare support (SBA 0.27 mmol g−1). PEI-SBA does adsorb more NO2 than MePEI-SBA, but it also has a lower amine efficiency (1.10 compared to 1.33). Given that this is a significantly higher uptake than both SO2 and CO2 (Table S2), it is likely that amine blockage and diffusion limitations play a more significant role in hindering CO2 and SO2 uptake. However, this observation may also be attributable to differences in amine−NO2 binding mechanism for different amine types.19,20 The PEI-SBA is likely to be the less restrictive of the two, as the favorable polymer−wall interactions will limit the “plug” morphology.15,19 In contrast, the MePEI-SBA may favor polymer−polymer interactions, due to the lack of primary amines, causing the polymer to “plug” in the pore, restricting diffusion through the polymer, and hindering access to the available amine sites.19,20 Again we note that the tertiary amines are able to adsorb NO2 and therefore could be useful as selective NO2 sorbents under “dry” conditions. The NO2 capacity of both MePEI-SBA and PEI-SBA decreases after oxidation at 90 °C (Figure 7A, Table S2) and decreases more rapidly once oxidized at 110 °C. The decrease is more pronounced in the PEI-SBA series. This agrees with our hypothesis that the amine sites in PEI-SBA are more available. If the NO2 uptake of MePEI-SBA is indeed hindered by diffusion due to its “plug” morphology, then loss of some of the MePEI (through cracking or evaporation) will make other amine sites accessible, causing less of a change in NO2 uptake. However, if the sites in PEI-SBA are generally more available, then removing any aminopolymer will only remove accessible, active amine species, causing a more rapid decay in NO2 uptake. Unlike the other acid gases tested, when the aminopolymers are oxidized at 130 °C, they still retain a significant amount of H

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels Present Address

adsorbed NO2, whereas the MePEI110-SBA and MePEI130SBA desorb 70−75% of the adsorbed NO2 (Table S3). This suggests that for the oxidatively degraded MePEI-SBA and PEI-SBA samples the majority of the NO2 species are reversibly adsorbed; however, higher temperatures (>110 °C) may be required to achieve quantitative desorption.

§

Matthew E. Potter is presently at the University of Southampton, Department of Chemistry, Southampton, Hants., SO17 1BJ, UK.

Funding



This work was supported by the Center for Understanding and Control of Acid Gas-Induced Evolution of Materials for Energy (UNCAGE-ME), an Energy Frontier Research Center, funded by U.S. Department of Energy (US DoE), Office of Science, Basic Energy Sciences (BES), under Award DE-SC0012577 and also by a Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2018R1A6A3A01012374).

CONCLUSION In this study we have demonstrated how extended oxidative degradation, mimicking aerobic thermal desorption conditions, can play a determining role in sorbent lifetime and uptake capacity of a range of acid gases. We have compared the oxidative degradation behavior of two encapsulated aminopolymers, varying only in their amine types, with analogous isolated grafted amines. We probed their capacity for adsorbing CO2, and common impurities such as SO2 and NO2, which have previously been shown to rapidly degrade supported amine sorbents. By controlling the distribution of primary, secondary and tertiary amines in both Class 1 and 2 sorbents, we are able to consider the behavior of each amine in the oxidation process. It was shown that CO2 and SO2 sorption are severely affected by oxidative degradation, with the resulting most oxidized supported amines losing all uptake capacity. In contrast, isolated Class 2 species were found to be notably more stable for adsorbing simulated dry flue gas (10 vol % CO2); however, with the exception of isolated tertiary amines (DMAPS-SBA), they were unable to adsorb significantly more SO2 than the bare SBA support after oxidation above 110 °C. Therefore, tertiary amines may be stable and selective SO2 sorbents, prior to oxidative degradation. In stark contrast, all systems were able to adsorb notable quantities of NO2, even once oxidized at 130 °C. This finding suggests that the oxidized products of these supported amines could show promise as a novel sorbent for the selective adsorption of NO2. This opens unique avenues for recycling used (oxidized) supported amine sorbents. Thus, once supported amines are unable to adsorb CO2 effectively they can instead be used to preremove NO2 from CO2 sources to preserve active supported amine sorbents for CO2 removal.



Notes

The authors declare no competing financial interest.



(1) Samanta, A.; Zhao, A.; Shimizu, G. K. H.; Sarkar, P.; Gupta, R. Post-Combustion CO2 Capture Using Solid Sorbents: A Review. Ind. Eng. Chem. Res. 2012, 51, 1438−1463. (2) Bonan, G. B. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science 2008, 320, 1444−1449. (3) von der Assen, N.; Müller, L. J.; Steingrube, A.; Voll, P.; Bardow, A. Selecting CO2 Sources for CO2 Utilization by EnvironmentalMerit-Order Curves. Environ. Sci. Technol. 2016, 50, 1093−1101. (4) Zou, L.; Sun, Y.; Che, S.; Yang, X.; Wang, X.; Bosch, M.; Wang, Q.; Li, H.; Smith, M.; Yuan, S.; Perry, Z.; Zhou, H. C. Porous Organic Polymers for Post-Combustion Carbon Capture. Adv. Mater. 2017, 29, 1700229−1700263. (5) Rao, A. B.; Rubin, E. S. A Technical, Economic, and Environmental Assessment of Amine-Based CO2 Capture Technology for Power Plant Greenhouse Gas Control. Environ. Sci. Technol. 2002, 36, 4467−4475. (6) Wang, Q.; Luo, J.; Zhong, Z.; Borgna, A. CO2 capture by solid adsorbents and their applications: current status and new trends. Energy Environ. Sci. 2011, 4, 42−55. (7) Serna-Guerrero, R. S.; Da’na, E.; Sayari, A. New Insights into the Interactions of CO2 with Amine-Functionalized Silica. Ind. Eng. Chem. Res. 2008, 47, 9406−9412. (8) Chen, C.; Ahn, W. S. CO2 capture using mesoporous alumina prepared by a sol−gel process. Chem. Eng. J. 2011, 166, 646−651. (9) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. An Amine-Functionalized MIL-53 Metal−Organic Framework with Large Separation Power for CO2 and CH4. J. Am. Chem. Soc. 2009, 131, 6326−6327. (10) Wang, J.; Huang, H.; Wang, M.; Yao, L.; Qiao, W.; Long, D.; Ling, L. Direct Capture of Low-Concentration CO2 on Mesoporous Carbon-Supported Solid Amine Adsorbents at Ambient Temperature. Ind. Eng. Chem. Res. 2015, 54, 5319−5327. (11) Chai, S. H.; Liu, Z. M.; Huang, K.; Tan, S.; Dai, S. Amine Functionalization of Microsized and Nanosized Mesoporous Carbons for Carbon Dioxide Capture. Ind. Eng. Chem. Res. 2016, 55, 7355− 7361. (12) Bae, T. H.; Hudson, M. R.; Mason, J. A.; Queen, W. L.; Dutton, J. J.; Sumida, K.; Micklash, K. J.; Kaye, S. S.; Brown, C. M.; Long, J. R. Evaluation of cation-exchanged zeolite adsorbents for post-combustion carbon dioxide capture. Energy Environ. Sci. 2013, 6, 128−138. (13) Fracaroli, A. M.; Furukawa, H.; Suzuki, M.; Dodd, M.; Okajima, S.; Gandara, F.; Reimer, J. A.; Yaghi, O. M. Metal−Organic Frameworks with Precisely Designed Interior for Carbon Dioxide Capture in the Presence of Water. J. Am. Chem. Soc. 2014, 136, 8863− 8866. (14) Lin, Y. C.; Kong, C. L.; Zhang, Q. J.; Chen, L. Metal-Organic Frameworks for Carbon Dioxide Capture and Methane Storage. Adv. Energy Mater. 2017, 7, 1601296.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.energyfuels.8b03688. Pore-size distributions and TGA analysis for all the samples used in this work, 1H NMR spectra of MePEI, and numerical values for CO2 (400 ppm and 10 vol %), SO2 (200 ppm), and NO2 (200 ppm) uptake with theoretical amine efficiencies (PDF)



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*Email: [email protected]. ORCID

Kyeong Min Cho: 0000-0001-8622-5451 Jason J. Lee: 0000-0002-1565-0706 Christopher W. Jones: 0000-0003-3255-5791 I

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

(36) Drese, J. H.; Choi, S.; Lively, R. P.; Koros, W. J.; Fauth, D. J.; Gray, M. L.; Jones, C. W. Synthesis−Structure−Property Relationships for Hyperbranched Aminosilica CO2 Adsorbents. Adv. Funct. Mater. 2009, 19, 3821−3832. (37) Wilfong, W. C.; Kail, B. W.; Jones, C. W.; Pacheco, C.; Gray, M. L. Spectroscopic Investigation of the Mechanisms Responsible for the Superior Stability of Hybrid Class 1/Class 2 CO2 Sorbents: A New Class 4 Category. ACS Appl. Mater. Interfaces 2016, 8, 12780− 12791. (38) Hahn, M. W.; Jelic, J.; Berger, E.; Reuter, K.; Jentys, A.; Lercher, J. A. Role of Amine Functionality for CO2 Chemisorption on Silica. J. Phys. Chem. B 2016, 120, 1988−1995. (39) Carrillo, J. M. Y.; Sakwa-Novak, M. A.; Holewinski, A.; Potter, M. E.; Rother, G.; Jones, C. W.; Sumpter, B. G. Unraveling the Dynamics of Aminopolymer/Silica Composites. Langmuir 2016, 32, 2617−2625. (40) Alkhabbaz, M. A.; Bollini, P.; Foo, G. S.; Sievers, C.; Jones, C. W. Important Roles of Enthalpic and Entropic Contributions to CO2 Capture from Simulated Flue Gas and Ambient Air Using Mesoporous Silica Grafted Amines. J. Am. Chem. Soc. 2014, 136, 13170−13173. (41) Bacsik, Z.; Ahlsten, N.; Ziadi, A.; Zhao, G.; Garcia-Bennett, A. E.; Martin-Matute, B.; Hedin, N. Mechanisms and Kinetics for Sorption of CO2 on Bicontinuous Mesoporous Silica Modified with nPropylamine. Langmuir 2011, 27, 11118−11128. (42) Vimont, A.; Lavalley, J. C.; Sahibed-Dine, A.; Arean, C. O.; Delgado, M. R.; Daturi, M. Infrared Spectroscopic Study on the Surface Properties of γ-Gallium Oxide as Compared to Those of γAlumina. J. Phys. Chem. B 2005, 109, 9656−9664. (43) Tailor, R.; Ahmadalinezhad, A.; Sayari, A. Selective removal of SO2 over tertiary amine-containing materials. Chem. Eng. J. 2014, 240, 462−468. (44) Lee, G. Y.; Lee, J.; Vo, H. T.; Kim, S.; Lee, H.; Park, T. AmineFunctionalized Covalent Organic Framework for Efficient SO2 Capture with High Reversibility. Sci. Rep. 2017, 7, 557−566. (45) Rezaei, F.; Rownaghi, A. A.; Monjezi, S.; Lively, R. P.; Jones, C. W. SOx/NOx Removal from Flue Gas Streams by Solid Adsorbents: A Review of Current Challenges and Future Directions. Energy Fuels 2015, 29, 5467−5486. (46) Bali, S.; Chen, T. T.; Chaikittisilp, W.; Jones, C. W. Oxidative Stability of Amino Polymer−Alumina Hybrid Adsorbents for Carbon Dioxide Capture. Energy Fuels 2013, 27, 1547−1554. (47) Heydari-Gorji, A.; Sayari, A. Thermal, Oxidative, and CO2Induced Degradation of Supported Polyethylenimine Adsorbents. Ind. Eng. Chem. Res. 2012, 51, 6887−6894. (48) Srikanth, C. S.; Chuang, S. S. C. Spectroscopic Investigation into Oxidative Degradation of Silica-Supported Amine Sorbents for CO2 Capture. ChemSusChem 2012, 5, 1435−1442. (49) Quang, D. V.; Hatton, T. A.; Abu-Zahra, M. R. M. Thermally Stable Amine-Grafted Adsorbent Prepared by Impregnating 3Aminopropyltriethoxysilane on Mesoporous Silica for CO2 Capture. Ind. Eng. Chem. Res. 2016, 55, 7842−7852. (50) Ahmadalinezhad, A.; Sayari, A. Oxidative degradation of silicasupported polyethylenimine for CO2 adsorption: insights into the nature of deactivated species. Phys. Chem. Chem. Phys. 2014, 16, 1529−1535. (51) Pang, S. H.; Lee, L. C.; Sakwa-Novak, M. A.; Lively, R. P.; Jones, C. W. Design of Aminopolymer Structure to Enhance Performance and Stability of CO2 Sorbents: Poly(propylenimine) vs Poly(ethylenimine). J. Am. Chem. Soc. 2017, 139, 3627−3630. (52) Sarazen, M. L.; Jones, C. W. Insights into Azetidine Polymerization for the Preparation of Poly(propylenimine)-Based CO2 Adsorbents. Macromolecules 2017, 50, 9135−9143. (53) Lukens, W. W.; Schmidt-Winkel, P.; Zhao, D.; Feng, J.; Stucky, G. D. Evaluating Pore Sizes in Mesoporous Materials: A Simplified Standard Adsorption Method and a Simplified Broekhoff−de Boer Method. Langmuir 1999, 15, 5403−5409.

(15) Potter, M. E.; Pang, S. H.; Jones, C. W. Adsorption Microcalorimetry of CO2 in Confined Aminopolymers. Langmuir 2017, 33, 117−124. (16) Sayari, A.; Heydari-Gorji, A.; Yang, Y. CO2-Induced Degradation of Amine-Containing Adsorbents: Reaction Products and Pathways. J. Am. Chem. Soc. 2012, 134, 13834−13842. (17) Dutcher, B.; Fan, M.; Russell, A. G. Amine-Based CO2 Capture Technology Development from the Beginning of 2013A Review. ACS Appl. Mater. Interfaces 2015, 7, 2137−2148. (18) Yan, X.; Zhang, L.; Zhang, Y.; Yang, G.; Yan, Z. AmineModified SBA-15: Effect of Pore Structure on the Performance for CO2 Capture. Ind. Eng. Chem. Res. 2011, 50, 3220−3226. (19) Holewinski, A.; Sakwa-Novak, M. A.; Jones, C. W. Linking CO2 Sorption Performance to Polymer Morphology in Aminopolymer/ Silica Composites through Neutron Scattering. J. Am. Chem. Soc. 2015, 137, 11749−11759. (20) Holewinski, A.; Sakwa-Novak, M. A.; Carrillo, J. M. Y.; Potter, M. E.; Ellebracht, N.; Rother, G.; Sumpter, B. G.; Jones, C. W. Aminopolymer Mobility and Support Interactions in Silica-PEI Composites for CO2 Capture Applications: A Quasielastic Neutron Scattering Study. J. Phys. Chem. B 2017, 121, 6721−6731. (21) Danon, A.; Stair, P. C.; Weitz, E. FTIR Study of CO2 Adsorption on Amine-Grafted SBA-15: Elucidation of Adsorbed Species. J. Phys. Chem. C 2011, 115, 11540−11549. (22) Hiyoshi, N.; Yogo, K.; Yashima, T. Adsorption characteristics of carbon dioxide on organically functionalized SBA-15. Microporous Mesoporous Mater. 2005, 84, 357−365. (23) Rezaei, F.; Jones, C. W. Stability of Supported Amine Adsorbents to SO2 and NOx in Postcombustion CO2 Capture. 1. Single-Component Adsorption. Ind. Eng. Chem. Res. 2013, 52, 12192−12201. (24) Tailor, R.; Abboud, M.; Sayari, A. Supported Polytertiary Amines: Highly Efficient and Selective SO2 Adsorbents. Environ. Sci. Technol. 2014, 48, 2025−2034. (25) Fan, Y.; Rezaei, F.; Labreche, Y.; Lively, R. P.; Koros, W. J.; Jones, C. W. Stability of amine-based hollow fiber CO2 adsorbents in the presence of NO and SO2. Fuel 2015, 160, 153−164. (26) Li, Y.; Buchi, S.; Grace, J. R.; Lim, C. SO2 Removal and CO2 Capture by Limestone Resulting from Calcination/Sulfation/Carbonation Cycles. Energy Fuels 2005, 19, 1927−1934. (27) Li, W.; Liu, Y.; Wang, L.; Gao, G. Using Ionic Liquid Mixtures To Improve the SO2 Absorption Performance in Flue Gas. Energy Fuels 2017, 31, 1771−1777. (28) Hajari, A.; Atanga, M.; Hartvigsen, J. L.; Rownaghi, A. A.; Rezaei, F. Combined Flue Gas Cleanup Process for Simultaneous Removal of SOx, NOx, and CO2A Techno-Economic Analysis. Energy Fuels 2017, 31, 4165−4172. (29) Son, W. J.; Choi, J. S.; Ahn, W. S. Adsorptive removal of carbon dioxide using polyethyleneimine-loaded mesoporous silica materials. Microporous Mesoporous Mater. 2008, 113, 31−40. (30) Srikanth, C. S.; Chuang, S. S. C. Infrared Study of Strongly and Weakly Adsorbed CO2 on Fresh and Oxidatively Degraded Amine Sorbents. J. Phys. Chem. C 2013, 117, 9196−9205. (31) Heydari-Gorji, A.; Belmabkhout, Y.; Sayari, A. Polyethylenimine-Impregnated Mesoporous Silica: Effect of Amine Loading and Surface Alkyl Chains on CO2 Adsorption. Langmuir 2011, 27, 12411−12416. (32) Zhao, W.; Zhang, Z.; Li, Z.; Cai, N. Investigation of Thermal Stability and Continuous CO2 Capture from Flue Gases with Supported Amine Sorbent. Ind. Eng. Chem. Res. 2013, 52, 2084−2093. (33) Bollini, P.; Choi, S.; Drese, J. H.; Jones, C. W. Oxidative Degradation of Aminosilica Adsorbents Relevant to Postcombustion CO2 Capture. Energy Fuels 2011, 25, 2416−2425. (34) Heydari-Gorji, A.; Belmabkhout, Y.; Sayari, A. Degradation of amine-supported CO2 adsorbents in the presence of oxygencontaining gases. Microporous Mesoporous Mater. 2011, 145, 146−149. (35) Khatri, R. A.; Chuang, S. S. C.; Soong, Y.; Gray, M. Thermal and Chemical Stability of Regenerable Solid Amine Sorbent for CO2 Capture. Energy Fuels 2006, 20, 1514−1520. J

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels (54) Zhai, Y.; Chuang, S. S. C. Enhancing Degradation Resistance of Polyethylenimine for CO2 Capture with Cross-Linked Poly(vinyl alcohol). Ind. Eng. Chem. Res. 2017, 56, 13766−13775.

K

DOI: 10.1021/acs.energyfuels.8b03688 Energy Fuels XXXX, XXX, XXX−XXX